Journal of South American Earth Sciences 91 (2019) 239–252
Contents lists available at ScienceDirect
Journal of South American Earth Sciences journal homepage: www.elsevier.com/locate/jsames
Rare earth element distribution in main lithologies of the Atibaia and Jaguari rivers’ subbasins (Southeast Brazil)
T
Bruno Cesar Mortatti, Luiza de Carvalho Mendes, Jacinta Enzweiler∗ Institute of Geosciences, University of Campinas/Unicamp, Rua Carlos Gomes 250, 13083-855, Campinas, SP, Brazil
A R T I C LE I N FO
A B S T R A C T
Keywords: Rare earth elements REE-bearing minerals Atibaia river Jaguari river Piracicaba basin MINSQ
In this study, we investigated the mineral and geochemical compositions of the main lithotypes constituting the Atibaia and Jaguari rivers’ subbasins (Southeast Brazil), with the aims of characterizing the distribution of rare earth elements (REE) in that diversified geological framework and identifying their source bearing minerals to the environment at the catchment scale. Samples from six geological units comprising granitoids of monzogabbroic to granitic compositions, orthogneisses (dioritic to granitic compositions), paragneisses and local occurrences of quartzites were studied. In estimates of mineral proportions in each analyzed lithotype, we identified plagioclase, K-feldspar, and Ca-amphibole followed by minor amounts of apatite, titanite, allanite and rarely monazite as the potential light REE hosting phases, while the dominant heavy REE carriers were Naamphibole, garnet, and zircon. The ∑REE contents in the rocks ranged between 59 and 791 mg kg−1, with a significant surplus of light over heavy REE [(La/Yb)CN = 11–197]. The whole-rock negative Eu anomalies of most samples result from the combined Eu budget of the individual minerals. Metamorphic lithotypes dominate at the Atibaia subbasin, whereas the main lithology of the Jaguari subbasin are granitoids. Comparatively, the first rocks contain slightly more ∑REE and are more enriched in light REE. These features and the relative weathering susceptibility of the REE hosting minerals constitute the initial step to explain the behavior of these elements in the environment.
1. Introduction The abundances of the rare earth elements (REE) and their relative fractionation serve as geochemical tracers for several geological processes. Igneous petrogenesis, rock weathering and pedogenesis, waterrock interactions and source signatures are fields that extensively make use of REE data (Nesbitt, 1979; Henderson, 1984; Taylor and McLennan, 1985, 1995; Goldstein and Jacobsen, 1988; Tricca et al., 1999; Aubert et al., 2001; Négrel, 2006; Steinmann and Stille, 2008; Hagedorn et al., 2011; Gaillardet et al., 2014; Armstrong-Altrin et al., 2015, 2017; Smith and Liu, 2018). The REE form a coherent group concerning chemical and physical properties. The trivalent oxidation state and lanthanide contraction along the series primarily control the distribution of REE in igneous systems (Goldschmidt, 1937). Additionally, Ce and Eu also occur naturally in tetravalent and bivalent states, respectively, which may result in the so-called anomalies in the normalized distribution patterns of the REE concentrations plotted against their respective atomic numbers. The normalizing reference can be a chondritic meteorite, a sedimentary “average rock” or an estimate of Earth's crust abundances
∗
(Henderson, 1984) or some other, depending on the sample matrix. Other useful practice consists in comparing the fractionation of the lighter elements of the group (LREE = La-Gd) to that of the heavier ones (HREE = Tb-Lu). Typically, the REE are trace constituents in geological matrices. Their relative abundances in igneous rocks depend on the composition of the source magma and physicochemical parameters that control the crystallization of minerals that can host the REE as trace elements or, less frequently, phases where these elements are essential constituents. During the rock cycle and, more specifically, chemical weathering, the REE behave as nonmobile elements (Gaillardet et al., 2014). The dissolved chemical species in pristine riverine waters depend strongly on the composition and weathering of the basement rocks together with runoff rates (Millot et al., 2002; Hartmann, 2009) and dry and wet deposition. In large watersheds such as the Amazon (5,900 × 103 km2 according to Warne et al., 2002), the extensive lithological variety results in diffuse inputs (Tardy et al., 2005) leading to a wide range of trace element concentrations in rivers (Gaillardet et al., 2014), including the REE. In stream waters from small catchments, the REE contents result from the local bedrock lithologies (Tricca et al.,
Corresponding author. E-mail address:
[email protected] (J. Enzweiler).
https://doi.org/10.1016/j.jsames.2019.01.017 Received 30 July 2018; Received in revised form 11 January 2019; Accepted 28 January 2019 Available online 05 February 2019 0895-9811/ © 2019 Elsevier Ltd. All rights reserved.
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
The Ribeira Belt comprises mainly Mesoproterozoic to Neoproterozoic supracrustal rocks of low metamorphic grade (Basei et al., 2008) reworked during the amalgamation of the Gondwana Supercontinent (700-470 Ma) (Tassinari et al., 2001). Three tectonic domains (São Roque, Embú and Costeiro) characterize its central segment (Tassinari et al., 2004). The São Roque Domain, represented by Serra do Itaberaba and São Roque groups, contains metasedimentary and metavolcanosedimentary rocks (Juliani and Beljavskis, 1995) intruded by syenogranitic, granodioritic and tonalitic granitoids of the Serra do Barro Branco Suite (Hackspacher et al., 1994; Sachs, 1999). Primarily comprising highly deformed and recrystallized supracrustal units of the Andrelândia and Itapira complexes (Ulbrich et al., 2005), the Alto Rio Grande Belt also comprehends migmatitic and orthogneissic rocks of the Amparo Complex (Peloggia, 1990). To the west, Paleozoic and Mesozoic sedimentary sequences of the Paraná Basin overlap those geotectonic compartments (Fig. 2). The area's largest geotectonic compartment is the Socorro-Guaxupé Nappe southern domain, which consists of granitic and metamorphic complexes (Vlach, 1985). The pluriserial Socorro Granitic Complex includes the Bragança Paulista, Salmão, Piracaia and Nazaré Paulista suites (Artur et al., 1993), whereas granulitic and gneissic rocks, mostly migmatized, form the majority of the metamorphic complexes (Paraisópolis, Piracaia, and Varginha) (Morais, 1999a). The Bragança Paulista Suite includes granitoids of monzodioritic, monzonitic, syenitic, granodioritic and granitic compositions (Campos Neto et al., 1984; Iwata, 1995) ranging from medium to high-K calcalkaline series (Artur et al., 1993). Intruding the Bragança Paulista Suite, the Salmão Suite consists of high-K calc-alkaline granitoids (Artur et al., 1993) with syenitic and granitic compositions (Campos Neto et al., 1984). The Piracaia Suite exhibits subalkaline affinity expressed by its quartz monzodiorites, hornblende-biotite quartz monzonites, monzodiorites and diorites (Gengo, 2014). According to the same author, granitoids and garnet-biotite granites with migmatitic features mainly form the Nazaré Paulista Suite. In the studied area, the Socorro Metamorphic Complex covers the Paraisópolis and the Piracaia complexes formed, respectively, by migmatites and charnockites (Braga, 2002) and schists, marbles, muscovite quartzites, gneisses with sillimanite, garnet and cordierite, hornblendebiotite gneisses, granitic and granodioritic gneisses, calc-silicate rocks and gondites (Morais, 1999b). Fig. 3 shows the lithological domains map of the Atibaia and Jaguari subbasins. The map displays the rock sampling sites identified by the herein adopted acronyms.
1999). Therefore, the hosting minerals and their weathering susceptibility constitute primary controls for the transference of the REE to the hydrosphere. In Brazil, except for the Amazon basin, the systematic study of the REE in surface waters is still limited. Campos and Enzweiler (2016) investigated the REE distributions in the Atibaia river waters (São Paulo State) and observed LREE enriched fractionation ratios, the opposite of that occurring for most of the world's rivers (Keasler and Loveland, 1982; Hoyle et al., 1984; Goldstein and Jacobsen, 1988; Elderfield et al., 1990; Tricca et al., 1999; Xu and Han, 2009). The increasing technological use of REE can also affect their concentrations in waterbodies of urbanized areas. One recurrent example is a positive Gd anomaly, also detected in Atibaia watershed water (Campos and Enzweiler, 2016). Atibaia river forms by the confluence of Cachoeira and Atibainha, after these watercourses leave their respective reservoirs. The latter two also receive water from the Jaguari-Jacareí dam, which is fed by the corresponding rivers. These interconnected reservoirs form the core of the Cantareira System that supplies water for approximately nine million inhabitants of the São Paulo metropolitan region (SABESP - Basic Sanitation Company of the State of São Paulo, 2017). The reservoirs' interconnection implies that Atibaia's river water composition is influenced by its watershed lithologies and by rocks of the Jaguari subbasin. Consequently, the understanding of the REE budget in its riverine waters requires a mineral and geochemical characterization of the lithotypes of the two subbasins. Previous studies (Campos Neto et al., 1984; Vlach, 1985; Artur et al., 1993; Juliani and Beljavskis, 1995; Iwata, 1995; Wernick et al., 1997; Ragatky, 1998; Janasi, 1999; Braga, 2002; Gengo, 2014) carried out for distinct purposes investigated geological aspects of areas within Atibaia and Jaguari subbasins. This study aimed to overview the geological framework of the two subbasins to distinguish them according to their lithological signatures by doing a comprehensive petrographic and chemical analysis of representative samples. In addition to assigning the relative REE distributions and fractionation patterns in the rocks of the area, we estimate the main hosting mineral proportions of these elements, departing from the thin section identified mineral assemblages together with major and selected trace elements data. 2. Study area The Atibaia and Jaguari subbasins constitute the northeastern portion of São Paulo State, aside from a small segment in the southwestern part of Minas Gerais State, both in Brazil. The two catchments cover an area of approximately 6000 km2 and are part of the Piracicaba drainage basin (11,403 km2) (CBH-PCJ - Committee of Hydrographic Basins of Piracicaba, 2016). The principal watercourses of the Atibaia subbasin include the Atibaia, Cachoeira and Atibainha rivers, where the last two form the Atibaia river. In the Jaguari subbasin, the Jaguari river along with its tributaries Camanducaia and Jacareí constitute the major water bodies (Fig. 1). The Jaguari, Jacareí, Cachoeira and Atibainha rivers feed four of the six reservoirs that form the Cantareira System, which is one of the most extensive water supply systems in the world, providing approximately 33 m3 s−1 of water to the São Paulo metropolitan region (SABESP - Basic Sanitation Company of the State of São Paulo, 2017).
3. Materials and methods To select the representative lithologies’ sampling sites from the Atibaia and Jaguari subbasins (Fig. 3), we used the CPRM (Brazilian Geological Survey) outcropping database allied to field observations. The twenty studied rock samples were collected from non-weathered outcrops preferentially at the vicinities of the major watercourses. Tables 1a and 1b list the sample identifications, the corresponding geological units, and subbasins as well as their respective field coordinates. Petrographic thin sections of the samples were used to describe the mineral composition with a microscope (ZEISS AxioPhot). For chemical analysis, representative rock subsamples were comminuted (< 75 μm) using a jaw crusher (Pulverisette 2 - Fritsch) and planetary and vibratory mills (respectively, Pulverisette 5 and 7 - Fritsch) with agate grinding devices. Major and minor element mass fractions were determined by X-ray fluorescence spectrometry (PW2404 - Philips) on glass discs obtained by fusion of pre-ignited powder test-portions (1.0 g) with lithium metaborate and tetraborate (6.0 g), using a method similar to that described by Vendemiatto and Enzweiler (2001). For the determination of the mass fractions of trace elements, test-portions (50 mg) were dissolved using a multi-acid digestion procedure in pressurized vessels,
2.1. Geological setting The study area is in the central sector of the Mantiqueira Province, a Neoproterozoic orogenic system in Southeast Brazil. The central portion of this province includes the Ribeira Orogen, the interference zone between Brasília and Ribeira orogens and Apiaí, São Roque and Embú terranes (Heilbron et al., 2004). Orogenic collapse events (590-550 Ma) during the Brasiliano Cycle structured the Ribeira Orogen (Machado et al., 1996; Heilbron and Machado, 2003) forming the Ribeira and Alto Rio Grande belts as well as the Socorro-Guaxupé Nappe (Fig. 2). 240
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 1. Limits of the Atibaia and Jaguari subbasins in the São Paulo (SP) and Minas Gerais (MG) states. 1 = Camanducaia river, 2 = Jaguari river, 3 = Jacareí river, 4 = Cachoeira river, 5 = Atibainha river, 6 = Atibaia river. JJR = Jaguari-Jacareí Reservoir, CR = Cachoeira Reservoir, AR = Atibainha Reservoir. Modified from CBH-PCJ (2016).
average measurement bias for the REE values in BRP-1 was less than 15%. The relative proportions of constituent minerals of each sample were quantified using a least squares spreadsheet (MINSQ) that was developed by Herrmann and Berry (2002) and employs the Solver tool from Microsoft Excel. To start, for each sample, we inserted the results of whole-rock analysis and the mineral assemblages recognized from petrographic descriptions. After several sequential steps, which consisted of including relevant trace element data and adding selected minerals into the MINSQ database, we arrived at a close approximation of the mineral composition of each lithotype, particularly concerning their potential REE-bearing phases. For the mineral data insertion, as reference criteria, we adopted the mineral formulas as they appear in the Webmineral database (www. webmineral.com). The intermediary members of some mineral groups such as eastonite (mica group - biotite subgroup), riebeckite (amphibole group), and almandine (garnet group) and continuous solid solution series such as plagioclase (albite-anorthite) were used in the MINSQ calculations, always supported by petrographic and chemical data.
followed by inductively coupled plasma mass spectrometry (ICP-MS, XSeries II - Thermo Fisher Scientific) analysis (Cotta and Enzweiler, 2012). The FeO mass fractions were determined using the method proposed by Amonette and Templeton (1998). Briefly, 50 mg test-portions were digested in acid-cleaned amber plastic bottles with 15 mL of an acid mixture containing 12 mL 10% (v/v) H2SO4, 2 mL 10% (m/v) 1,10phenanthroline monohydrate solution and 1 mL 48% HF at 100 °C for 30 min under agitation. The digest was neutralized with 10 mL of 5% H3BO3 (m/v) and brought up to 100 mL with ultrapure water. Aliquots of 1 mL mixed with 10 mL of a 1% sodium citrate solution were used to measure the absorbance at 510 nm on a spectrophotometer (UV–Vis B382, Micronal), after calibration with appropriate standard solutions of ferrous ethylenediammonium sulfate. For analytical quality control purposes, duplicates of selected samples and rock reference materials (GS-N, AC-E, and BRP-1 for XRF and GS-N and BRP-1 for ICP-MS) were analyzed. The measurement uncertainties for major and minor elements vary from 1.5% (SiO2) to 7% (P2O5) (Enzweiler and Vendemiatto, 2013), and, for trace elements, they ranged between 8 and 16%, all at the 95% confidence interval. The measurement bias obtained for the FeO mass fractions in granite reference materials AC-E and GS-N was less than 7%. In situ mineral composition data obtained on separate plagioclase and garnet grains of two samples were prepared as polished resin mounts and analyzed by laser ablation (Excite 193, Photon Machines) and sector field inductively coupled plasma mass spectrometry (Element XR, Thermo Fisher Scientific) (LA-SF-ICP-MS). The laser spot diameter was 85 μm, and the remaining conditions followed Navarro et al. (2015), including data processing (Iolite, version 2.4). The LA-ICP-MS was calibrated using NIST SRM 612 reference values (Jochum et al., 2011), and for analytical quality control, we used BRP-1G, a glass of the basalt reference material BRP-1 (Cotta and Enzweiler, 2008). The
4. Results 4.1. Rock mineral composition Tables 1a and 1b present the mineral compositions of the metamorphic and igneous rock samples, respectively. For each sample, the third columns contain the mineral assemblages identified from petrographic thin sections (numerator) and those calculated with the MINSQ spreadsheet (denominator). The residual sum of squares (RSSQ) expresses the difference between the actual (analyzed) and estimated (calculated) percent mass fractions of principal constituents and selected trace elements of each rock sample according to Herrmann and 241
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 2. Major geotectonic compartments in the studied region (modified from Morais, 1999a).
accessories. The Bragança Paulista Suite (BPS) is represented mainly by granitoids (granodioritic to granitic compositions) and to a lesser degree quartz monzodioritic rocks. The samples exhibited K-feldspar, plagioclase, quartz, biotite, amphibole and, less recurrently, muscovite followed by minor amounts of opaque minerals, apatite, zircon, and rarely epidote, titanite, chlorite, garnet, and monazite. The Piracaia Suite (PS) rocks include (quartz) diorites, (quartz) monzodiorites, (quartz) monzonites, (quartz) syenites, alkali-syenites, and granites (Wernick et al., 1997). In our dataset, the monzonite sample CAC1 represents the PS. In thin sections, we identified plagioclase, K-feldspar, biotite, amphibole, pyroxene, and subordinate quartz, whereas opaque minerals, tourmaline, and zircon were the main accessories phases (Table 1b). Artur et al. (1993) described a similar mineral assemblage for rocks of the PS and included titanite and apatite as accessories.
Berry (2002), who proposed that an RSSQ value of less than 0.5 is usually adequate. The last columns of Tables 1a and 1b provide the lithotype denomination. 4.1.1. Major and accessory mineral phases The lithotypes of the six principal geological units of the ASB and JSB comprise granitoids of monzogabbroic to granitic compositions, orthogneisses (dioritic to granitic compositions), paragneisses and local occurrences of quartzites. Samples of gneissic rocks of dioritic to granitic compositions, paragneisses, and a quartzite represent the Piracaia Complex (PC). Recurrent mineral phases include quartz, plagioclase, K-feldspar, biotite, and amphibole. Muscovite, cordierite, and pyroxene were also identified, but less frequently. Opaque minerals, zircon, garnet, apatite, allanite, titanite, rutile, and graphite, could be recognized among the accessory minerals (Table 1a). The Serra do Itaberaba Group (SIG) constitutes a restrict domain in the subbasins’ southeastern area (Fig. 3) and is represented by the quartzite (PIR1), which contains plagioclase, muscovite, biotite, and Kfeldspar, besides opaque minerals and zircon as accessories (Table 1a). The Nazaré Paulista Suite (NPS) samples are granites with biotite and amphibole. Minor amounts of sillimanite, garnet and muscovite were also described (Table 1b), in agreement with previous works (Janasi, 1999; Gengo, 2014). Samples ATN2 and PIR2 pertain to the Serra do Barro Branco Suite (SBBS) and comprise igneous rocks of granodioritic and granitic composition, respectively. Quartz, plagioclase, K-feldspar, biotite, amphibole and muscovite are the dominant minerals, followed by minor contents of opaque minerals, garnet, and zircon (Table 1b). Studying the same suite, Campos Neto et al. (1984) described plagioclase, microcline, quartz, hornblende, and biotite as common minerals, whereas titanite, zircon, apatite and opaque minerals were the most frequent
4.2. Geochemical characterization The mass fractions of major, minor and selected trace elements, the rock samples and their respective molar ratios [Al2O3/(Na2O + K2O)] (A/NK) and [Al2O3/(CaO + Na2O + K2O)] (A/CNK) or Shand indexes (Shand, 1943) are given in Table 2. The chemical classification of the total alkali vs. SiO2 (Middlemost, 1994, Fig. 4A) shows that most of our igneous lithotypes fall within the granite field, fewer are quartz-monzonites, while only one sample pertains to the syenite and another to the monzogabbro fields. The diagram of K2O vs. SiO2 (Peccerillo and Taylor, 1976, Fig. 4B) indicates that same samples present a predominant shoshonitic to a high-K calc-alkaline composition. The Shand's diagram (Fig. 4C) of all the samples shows that they split almost in the same number within the metaluminous and peraluminous fields, attesting to the geochemical compositional variability of the area. 242
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 3. Lithological domains map of the studied area (modified from Morais, 1999a, 1999b; 1999c; Sachs, 1999; CBH-PCJ - Committee of Hydrographic Basins of Piracicaba, 2016) with the rock sampling sites as identified in Tables 1a and 1b. Acronyms: SC = Sedimentary Coverage, PB = Paraná Basin, SBBS = Serra do Barro Branco Suite, SRG = São Roque Group, SIG = Serra do Itaberaba Group, BPS = Bragança Paulista Suite, SS = Salmão Suite, AS = Atibaia Suite, NPS = Nazaré Paulista Suite, PS = Piracaia Suite, JC = Jaguariúna Complex, MC = Morungaba Complex, PsC = Paraisópolis Complex, PC = Piracaia Complex, IR = Igneous Rocks, AC = Amparo Complex, IAGs = Andrelândia/Itapira Groups.
Table 1a Metamorphic rock samples identification (ID), geological unit (GU), subbasin (SB), field coordinates, mineral assemblages determined from petrographic analysis and estimated by MINSQ with their relative difference (RSSQ) and respective classification. Sample ID a GU/bSB
UTM X UTM Y
ATB1 PC/ASB ATB4 PC/ASB
297097 E 7465481 S 336955 E 7444395 S
ATB5 PC/ASB
339995 E 7444612 S
BEC2 PC/JSB
359580 E 7469474 S
BEC4 PC/JSB
381343 E 7482944 S
BEC5 PC/JSB
383801 E 7476163 S
CAC1.5 PC/ASB
378830 E 7461259 S
PIR1 SIG/ASB
372352 E 7446815 S
a b c
Minerals in thin section MINSQ calculated mineral proportions (wt.%) [RSSQ = Residual Sum of Squares]
Rock classification
Qz + Ms ± Opq (Gr) Qz (75) + Ms (21) + Phl (1) + Mag (1) + Gr (0.5) + Rt (0.2) [RSSQ = 0.3] Pl + Qz + Kfs + Bt ± Opq ± Grt ± Ap ± Zrn Pl (42 ≈ Ab64An36) + Qz (30) + Eas (9) + Kfs (9) + Alm (4) + Mag (2) + Ilm (2) + Ap (1) + Py (0.2) + Zrn (0.03) [RSSQ = 0.1] Qz + Pl + Kfs + Crd + Cam + Bt ± Opq ± Grt ± Zrn Qz (48) + Pl (15 ≈ Ab67An33) + Crd (12) + Kfs (10) + Ann (7) + Mg-Hbl (4) + Mag (2) + Ilm (1) + Sps (0.4) + Py (0.05) + Zrn (0.02) [RSSQ = 0.0] Qz + Kfs + Pl + Ms + Bt + Cam + Cpx ± Grt ± Zrn Pl (26 ≈ Ab81An19) + Qz (24) + Kfs (23) + Ms (8) + Mg-Hbl (7) + Ann (6) + Aeg (4) + Rt (0.8) + Ap (0.7) + Sps (0.2) + Py (0.08) + Zrn (0.07) [RSSQ = 0.0] Pl + Qz + Bt + Kfs + Cam ± Opq ± Grt ± Zrn Pl (56 ≈ Ab64An36) + Qz (13) + Kfs (10) + Cum (8) + Ann (7) + Mag (3) + Rt (1) + Ap (0.6) + Sps (0.4) + Zrn (0.03) + Py (0.01) [RSSQ = 0.1] Kfs + Qz + Pl + Bt + Cam ± Opq ± Grt ± Rt ± Aln ± Zrn Kfs (36) + Qz (31) + Pl (28 ≈ Ab82An18) + Mag (3) + Eas (1) + Fe-Hbl (1) + Rt (0.4) + Ap (0.2) + Zrn (0.07) + Sps (0.05) + Aln (0.04) [RSSQ = 0.0] Kfs + Qz + Pl + Bt + Ms ± Opq ± Ttn ± Ap ± Aln ± Zrn Qz (27) + Kfs (25) + Pl (23 ≈ Ab83An17) + Ann (9) + Eas (7) + Ttn (4) + aluminoceladonite (2) + Ap (2) + Mag (0.1) + Aln (0.1) + Zrn (0.06) [RSSQ = 0.1] Qz + Pl + Ms + Bt + Kfs ± Opq ± Zrn Qz (72) + Pl (17 ≈ Ab88An12) + Ms (4) + Eas (2) + Kfs (2) + Ann (1) + Mag (0.7) + Rt (0.2) + Lzl (0.1) + Zrn (0.03) [RSSQ = 0.0]
graphite-bearing muscovite quartzite garnet-bearing biotite orthogneiss
c
garnet-bearing biotitehornblende-cordierite paragneiss garnet-bearing clinopyroxenehornblende-biotite orthogneiss garnet-bearing hornblendebiotite orthogneiss allanite-garnet-bearing hornblende-biotite granitic gneiss allanite-bearing muscovitebiotite paragneiss biotite-muscovite-plagioclase quartzite
GU = Geological Unit being NPS = Nazaré Paulista Suite, SBBS = Serra do Barro Branco Suite, BPS = Bragança Paulista Suite, and PS = Piracaia Suite. SB = Subbasin being ASB = Atibaia Subbasin and JSB = Jaguari Subbasin. Estimates obtained by MINSQ trial solution (see text). The used mineral symbols followed Whitney and Evans (2010) recommendations. 243
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Table 1b Igneous rock samples identification (ID), geological unit (GU), subbasin (SB), field coordinates, mineral assemblages determined from petrographic analysis and estimated by MINSQ with their relative difference (RSSQ) and respective classification. Sample ID a GU/bSB
UTM X UTM Y
NBJ1 NPS/ASB
351636 E 7438088 S
NBJ4 NPS/ASB
351636 E 7438088 S
ATN2 SBBS/ASB
363222 E 7434831 S
PIR2 SBBS/ASB
371762 E 7445170 S
ATB2 BPS/ASB
300462 E 7464231 S
ATB3 BPS/ASB ATB6 BPS/ASB
318665 E 7454827 S 343507 E 7437030 S
BEC1 BPS/JSB
352889 E 7464742 S
BEC3 BPS/JSB
371702 E 7478878 S
CAC0 BPS/ASB
382332 E 7458486 S
CAC2 BPS/ASB
379720 E 7459679 S
CAC1 PS/ASB
356076 E 7446549 S
Minerals in thin section MINSQ calculated mineral proportions (wt.%) [RSSQ = Residual Sum of Squares]
d
Pl + Qz + Kfs + Cam + Bt ± Opq ± Sil ± Ap ± Ms ± Grt ± Zrn Pl (36 ≈ Ab86An14) + Qz (31) + Kfs (26) + Eas (2) + Ann (1) + Rbk (1) + Sil (1) + Ilm (0.3) + Ap (0.2) + Ms (0.06) + Sps (0.02) + Zrn (0.02) + Py (0.003) [RSSQ = 0.0] Pl + Qz + Kfs + Bt ± Opq ± Sil ± Grt ± Ms ± Zrn Pl (38 ≈ Ab87An13) + Qz (32) + Kfs (25) + Eas (1) + Sil (1) + Alm (1) + Mag (0.7) + Ap (0.2) + Rt (0.1) + Zrn (0.02) + Ms (0.01) [RSSQ = 0.0] Pl + Qz + Kfs + Cam + Bt ± Opq ± Grt ± Zrn Pl (42 ≈ Ab98An2) + Qz (20) + Kfs (16) + Eas (9) + Fe-Hbl (8) + Adr (3) + Ap (0.6) + Mag (0.6) + Rt (0.6) + Zrn (0.04) + Py (0.04) [RSSQ = 0.0] Kfs + Pl + Qz + Bt + Ms ± Opq ± Grt ± Zrn Pl (32 ≈ Ab91An9) + Kfs (30) + Qz (26) + Alm (4) + Eas (3) + Ms (3) + Mag (1) + Rt (0.5) + Ap (0.4) + Zrn (0.05) [RSSQ = 0.0] Qz + Pl + Kfs ± Ep ± Chl ± Ms ± Opq ± Mnz Pl (37 ≈ Ab100An0) + Qz (34) + Kfs (26) + Ep (0.7) + Mg-Chl (0.4) + Ms (0.4) + Ilm (0.3) + Mnz (0.04) [RSSQ = 0.0] Qz + Pl + Kfs + Bt ± Grt ± Ep ± Opq Pl (33 ≈ Ab73An27) + Qz (31) + Kfs (26) + Eas (4) + Alm (3) + Ep (2) + Ilm (0.5) [RSSQ = 0.1] Kfs + Qz + Pl + Bt + Ms + Cam ± Opq ± Ap ± Zrn Kfs (34) + Qz (30) + Pl (28 ≈ Ab93An7) + Ms (3) + Ann (2) + Mag (1) + Mg-Hbl (1) + Ilm (0.5) + Ap (0.1) + Zrn (0.05) [RSSQ = 0.0] Pl + Kfs + Qz + Bt + Cam ± Opq ± Ttn ± Ap ± Zrn Pl (31 ≈ Ab84An16) + Kfs (22) + Qz (22) + Eas (10) + Fe-Hbl (8) + Ttn (4) + Mag (2) + Ap (1) + Zrn (0.04) [RSSQ = 0.2] Pl + Cam + Bt + Kfs + Qz ± Opq ± Ap ± Zrn Pl (38 ≈ Ab47An53) + Rbk (15) + Prg (14) + Eas (12) + Kfs (9) + Qz (6) + Mag (2) + Rt (2) + Ap (1) + Zrn (0.06) [RSSQ = 0.0] Kfs + Pl + Qz + Bt + Cam ± Opq ± Ap ± Zrn Pl (32 ≈ Ab78An22) + Kfs (32) + Qz (23) + Mag (5) + Eas (5) + Ap (1) + Fe-Hbl (1) + Rt (0.9) + Zrn (0.08) [RSSQ = 0.1] Kfs + Qz + Pl + Ms + Cam ± Opq ± Bt Kfs (32) + Qz (32) + Pl (26 ≈ Ab92An8) + Ms (5) + Mg-Hbl (2) + Mag (0.8) + Ann (0.3) + Ilm (0.3) [RSSQ = 0.1] Pl + Kfs + Bt + Cam + Cpx + Qz ± Opq ± Tur ± Zrn Pl (38 ≈ Ab95An5) + Kfs (35) + siderophyllite (8) + Mg-Hbl (6) + Aeg (4) + Qz (3) + Srl (2) + Ilm (2) + Ap (0.5) + Zrn (0.08) + Py (0.03) [RSSQ = 0.1]
garnet-muscovite-sillimanitebearing biotite-hornblende monzogranite muscovite-garnet-sillimanitebearing biotite monzogranite
Rock classification
c
garnet-bearing biotitehornblende granodiorite garnet-bearing muscovitebiotite monzogranite muscovite-bearing monzogranite garnet-bearing biotite monzogranite hornblende-muscovite-biotite monzogranite hornblende-biotite granodiorite biotite-hornblende quartz monzodiorite hornblende-biotite monzogranite hornblende-muscovite monzogranite tourmaline-bearing clinopyroxene-hornblendebiotite monzonite
a
GU = Geological Unit being NPS = Nazaré Paulista Suite, SBBS = Serra do Barro Branco Suite, BPS = Bragança Paulista Suite, and PS = Piracaia Suite. SB = Subbasin being ASB = Atibaia Subbasin and JSB = Jaguari Subbasin. c Estimates obtained by MINSQ trial solution (see text). d Rocks were classified according to the normative QAPF diagram (Streckeisen, 1974). The used mineral symbols followed Whitney and Evans (2010) recommendations. b
adopted in this work (see Table 3 footnotes). The LREE enrichment of PC lithotypes is noticeable in the chondritenormalized REE diagram (Fig. 5A) and partially reflects the predominance of plagioclase as well as K-feldspar in the analyzed samples (Table 1a), but not only. Two gneisses contain allanite, an LREE bearing phase, identified in thin sections and estimated by MINSQ (BEC5 = 0.04 wt.% and CAC1.5 = 0.1 wt.%). According to our calculated mineral compositions (Table 1a), the granitic gneiss BEC5, the paragneiss (ATB5) and the orthogneiss (BEC2) also contain Ca-amphiboles, a phase slightly enriched in LREE compared to HREE (Bea, 1996). The average HREE value in PC rock samples (13 mg kg−1) can be attributed to garnet and zircon, as also documented in other studies (Tapia-Fernandez et al., 2017; Armstrong-Altrin et al., 2018). We estimated 4 wt.% almandine in orthogneiss ATB4, whereas the orthogneiss BEC4 and the paragneiss ATB5 contain spessartine (both 0.4 wt.%) followed by lower contents of zircon (Table 1a). The REE mass fractions in separated plagioclase and garnet grains from orthogneiss ATB4 obtained by LA-SF-ICP-MS are listed in Table A2 (Appendix A - Supplementary Data), and the respective chondritenormalized distribution patterns are displayed in Figs. 6 and 7, respectively. The chondrite-normalized REE distribution patterns of the analyzed plagioclase grains from orthogneiss ATB4 (Fig. 6) and the respective whole-rock (Fig. 5A) are similar, consistent with the lithotype's plagioclase content (> 40 wt.%) (Table 1a). On the other hand, data from
Fig. 4A to C show the diagrams used to classify igneous and, when appropriate, metamorphic rock samples (e.g., Shand's diagram in Fig. 4C). Table 3 presents the REE and Y mass fractions, along with totals (∑REE, ∑LREE, and ∑HREE), the LREE/HREE and (La/Yb)CN ratios, where CN stands for chondrite normalized (Sun and McDonough, 1989), as well as the Eu anomalies values (Eu/Eu*), calculated according McLennan (1989). The REECN distribution patterns (Fig. 5A to D) show that all samples are enriched in LREE compared to the HREE. In these diagrams, metamorphic rock samples are displayed in Fig. 5A, while the igneous lithotypes are shown in Fig. 5B to D, mostly assembled by geological units. 4.2.1. Piracaia Complex The metamorphic rock samples of PC contain relatively wide compositional ranges of SiO2 (59.48–85.31 wt.%), Al2O3 (7.95–17.23 wt. %), Fe2O3 (0.16–2.36 wt.%) and FeO (0.34–4.12 wt.%) (Table 2), which correspond to gneisses (ortho and para) and one quartzite. The REE content of these samples also showed a large interval (∑REE = 123–791 mg kg−1), enrichment in LREE (LREE/ HREE = 13–71) and moderate fractionation [(La/Yb)CN = 11–87] (Table 3). The above REE totals are higher than those reported for similar rocks of the same complex (∑REE = 89–313 mg kg−1, LREE/ HREE = 19–36, [La/Yb]CN = 16–50) (Ragatky, 1998), where the LREE/HREE were recalculated according to the lanthanides division 244
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Table 2 Major, minor and selected trace element mass fractions (oxides in % and trace elements in mg kg−1) for the analyzed rock samples of the Atibaia and Jaguari subbasins. Sample
ATB1
Unit
a
SiO2 TiO2 Al2O3 Fe2O3 FeO MnO MgO CaO Na2O K2O P2O5 LOI* Total A/NK A/CNK Ba Hf Nb Rb Sr Ta Th Zr
85.31 0.165 7.95 0.91 0.34 0.008 0.27 0.01 0.11 3.10 0.060 1.37 99.6 2.25 2.24 696 2.81 2.34 104 59.2 0.42 15.2 98.6
ATB4
ATB5
BEC2
BEC4
BEC5
CAC1.5
PC
PIR1
NBJ1
b
c
86.80 0.219 6.83 0.50 0.80 0.021 0.47 0.63 1.68 1.16 0.063 0.77 99.9 1.70 1.32 354 5.25 4.92 24.3 81.2 0.27 6.00 217
73.57 0.159 14.40 0.20 0.85 0.010 0.34 1.40 3.59 4.75 0.074 0.48 99.8 1.30 1.02 542 3.37 7.33 140 200 0.28 11.4 115
SIG
64.41 0.765 16.60 1.37 3.27 0.113 1.84 4.14 3.24 2.55 0.342 0.84 99.5 2.10 1.06 733 6.21 12.5 125 457 0.77 4.30 207
74.19 0.527 11.05 1.39 3.91 0.170 2.46 1.72 1.12 2.29 0.054 0.94 99.8 2.58 1.48 629 3.68 7.44 108 118 0.45 28.6 119
66.43 0.766 14.82 1.42 2.47 0.078 1.28 2.57 2.90 5.32 0.316 1.06 99.4 1.45 0.98 2063 9.09 20.8 208 598 0.85 11.8 402
59.48 1.169 17.23 2.36 3.99 0.144 2.76 4.54 4.01 2.21 0.420 1.23 99.5 1.97 1.00 750 4.03 21.7 84.7 485 0.61 1.33 176
72.31 0.382 13.24 2.00 1.15 0.025 0.19 1.34 2.56 6.17 0.066 0.25 99.7 1.24 0.99 1136 10.3 6.67 129 130 0.10 10.9 435
65.99 1.137 13.95 0.16 4.12 0.063 1.50 2.99 2.10 6.18 0.701 1.08 100.0 1.42 0.89 2358 9.56 26.6 319 382 1.98 25.1 375
NBJ4
ATN2
PIR2
d
NPS
65.24 0.582 15.57 1.65 2.60 0.071 1.65 3.09 4.57 3.71 0.277 0.70 99.7 1.40 0.91 1629 7.09 11.9 131 933 0.82 10.6 256
ATB3
ATB6
BEC1
BEC3
CAC0
CAC2
e
SBBS
73.35 0.098 14.49 0.51 0.70 0.019 0.23 1.44 3.76 4.29 0.085 0.35 99.3 1.30 1.08 1047 3.81 5.02 128 273 0.17 25.8 133
ATB2
f
BPS
69.46 0.503 15.53 0.51 1.87 0.024 0.64 1.02 3.22 5.86 0.162 0.99 99.8 1.36 1.15 1835 7.92 5.01 156 646 0.14 23.3 328
76.62 0.108 12.59 0.33 0.25 0.051 0.08 0.53 4.10 4.52 0.022 0.29 99.5 1.15 1.00 157 2.76 11.7 112 36.4 0.80 13.1 63.8
CAC1
70.63 0.293 14.91 0.75 1.49 0.071 0.73 2.35 2.68 4.75 0.182 0.57 99.4 1.58 1.08 1533 4.34 11.6 197 486 1.29 26.0 130
72.71 0.266 13.41 0.79 1.27 0.037 0.18 0.79 2.92 6.23 0.049 0.61 99.3 1.15 1.03 622 5.96 17.6 184 70.5 0.92 14.1 219
63.23 1.017 15.52 1.84 3.13 0.076 1.97 4.24 3.13 4.88 0.343 0.62 100.0 1.52 0.86 2315 6.04 20.7 104 942 1.03 0.94 249
49.44 2.000 19.44 4.06 5.23 0.139 4.34 6.61 3.63 2.86 0.591 1.05 99.4 2.15 0.93 1051 8.04 38.0 160 756 2.65 6.94 344
65.11 0.911 15.26 3.53 1.85 0.078 1.04 2.55 2.92 5.99 0.355 0.62 100.2 1.39 0.94 1628 11.1 24.0 195 279 1.12 5.20 487
73.62 0.134 13.78 0.58 0.50 0.012 0.37 0.86 2.72 6.13 0.030 0.51 99.3 1.24 1.09 2407 1.50 9.79 240 373 0.59 6.79 40.2
PS
58.78 0.843 18.53 1.43 3.86 0.124 1.11 1.78 4.62 6.78 0.201 0.57 99.9 1.24 1.02 3102 11.2 37.3 145 479 1.70 5.50 470
a
Piracaia Complex. Serra do Itaberaba Group. c Nazaré Paulista Suite. d Serra do Barro Branco Suite. e Bragança Paulista Suite. f Piracaia Suite, LOI* = Loss on ignition (1000 °C) corrected according to the following formula: LOI* = LOIobtained + (FeOFe2O3 - FeO), A/NK = molar ratio of Al2O3/(Na2O + K2O), A/CNK = molar ratio of Al2O3/(CaO + Na2O + K2O) (Shand, 1943). b
Yb)CN = 72–94] (Table 3). Janasi (1999) observed similar fractionation in samples of the same geological unit. The REECN diagram (Fig. 5B) illustrates the distribution patterns of both samples, which are almost parallel, but with higher REE mass fractions in NBJ4 compared to NBJ1. This feature may derive from distinct crust melting rates or larger differentiation during NBJ4 rock genesis compared to NBJ1 or even due to merely compositional heterogeneity between the analyzed samples. We were not able to associate the observed differences to any particular petrographic aspect given the estimated mineral assemblages; also, the relative proportion of main REE carriers are quite similar for both samples (Table 1b). The NPS granite samples exhibited negative Eu anomalies (Eu/ Eu* = 0.5 for NBJ1 and 0.4 for NBJ4) and a simultaneous Sr depletion in the chondrite-normalized multi-element diagram (Fig. A1; Appendix A), suggesting the fractional crystallization of plagioclase. Ragatky (1998) presented similar values (Eu/Eu* = 0.4), while Janasi (1999) reports less prominent negative Eu anomalies (Eu/Eu* = 0.6–0.8) for samples of the NPS. Fig. 8 shows the chondrite-normalized pattern of the REE mass fractions in plagioclase grains from NBJ1 (Table A2 - Appendix A). The plagioclase grains presented an expressive Eu positive anomaly (Eu/ Eu* = 6.8; Table A2), while the whole-rock has a negative one (Eu/ Eu* = 0.5; Table 3). In this context, accessory phases such as apatite, garnet, and zircon, which are usually strongly depleted in Eu, seem to be dominating the observed whole-rock negative Eu anomaly.
garnet grains of the same sample (Fig. 7) exhibited HREE enriched chondrite-normalized REE patterns and negative Eu anomaly (Table A2 in Appendix A for Eu/Eu* values). The rocks of this geological unit are characterized by negative Eu anomalies (Eu/Eu* = 0.4–0.8; Table 3), which we attribute to the contribution of accessory minerals (garnet, zircon, and allanite, among others) to the analyzed rocks REE budget. Ragatky (1998) reported a similar range of negative Eu anomalies (Eu/Eu* = 0.4–1.1) for the PC rocks. 4.2.2. Serra do Itaberaba Group The biotite-muscovite-plagioclase quartzite (PIR1) with its composition rich in SiO2 (86.80 wt.%), low Al2O3 (6.83 wt.%) pertains to the peraluminous field based on its Shand's indexes (Fig. 4C). PIR1 has low REE content (∑REE = 92 mg kg−1; Table 3) and small LREE enrichment (LREE/HREE = 14) and fractionation [(La/Yb)CN = 14] compared to most of this study's rock samples (Table 3). According to the petrographic description and MINSQ mineral proportion, the only significant HREE bearing phase in PIR1 is zircon (≈0.03 wt.%; Table 1a). PIR1 presented negative Eu and Ce anomalies (Eu/Eu* = 0.7; Table 3 and Fig. 5A). This last feature may be a result of Ce oxidation and removal during early diagenetic processes (Pattan et al., 2005). 4.2.3. Nazaré Paulista Suite The NPS monzogranite samples (NBJ1 and NBJ4; Fig. 4A and Table 2) exhibit predominant high-K calc-alkaline composition (Fig. 4B) and peraluminous characteristics (Fig. 4C) according to the Shand's indexes (A/NK ≈ 1.30 and A/CNK ≈ 1.02 and 1.08, respectively) caused mainly by their garnet and sillimanite contents (Table 1b). NBJ1 and NBJ4 contain intermediary REE mass fractions of 119 and 254 mg kg−1, respectively and, at the same time, prominent LREE enrichment (LREE/HREE = 41–50) and high fractionation ratios [(La/
4.2.4. Serra do Barro Branco Suite The granodiorite (ATN2) and monzogranite (PIR2) fell in the high-K calc-alkaline and shoshonitic compositional fields, respectively (Fig. 4B), while their Shand's indexes (Fig. 4C, Table 2) indicate metaluminous (ATN2) and peraluminous (PIR2) characteristics. ATN2 and PIR2 contain intermediate to high ∑REE (255 and 245
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 4. A, B and C - Geochemical characteristics of the studied samples identified by their geological units. A) Na2O + K2O vs. SiO2 classification diagram (Middlemost, 1994). B) K2O vs. SiO2 magmatic series composition diagram (Peccerillo and Taylor, 1976). C) A/NK vs. A/CNK Shand's diagram (Shand, 1943).
246
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Table 3 REEY mass fractions (mg kg−1) for the analyzed rock samples of the Atibaia and Jaguari subbasins. Sample
ATB1
Unit
a
Y La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu ∑REE ∑LREE ∑HREE LREE/HREE (La/Yb)CN Eu/Eu* ∑REEg ∑LREEg ∑HREEg
12.2 27.1 54.1 5.82 22.1 4.08 0.69 3.25 0.46 2.41 0.43 1.22 0.16 1.01 0.14 123 117 5.83 20 19 0.6 286 271 11
a b c d e f g
ATB4
ATB5
BEC2
BEC4
BEC5
CAC1.5
PC 25.5 36.3 71.5 8.20 33.8 6.37 1.67 5.93 0.94 4.94 0.89 2.61 0.33 2.27 0.33 176 164 12.31 13 11 0.8
36.8 69.7 137 15.5 60.7 10.2 1.38 8.73 1.16 6.34 1.30 4.03 0.54 3.81 0.54 321 303 17.72 17 13 0.4
39.0 107 213 24.0 86.0 14.2 2.80 11.0 1.43 7.72 1.43 3.69 0.48 3.03 0.43 476 458 18.21 25 25 0.7
24.9 58.5 112 11.8 42.1 6.90 1.65 6.14 0.85 4.86 0.94 2.35 0.28 1.62 0.24 250 239 11.14 21 26 0.8
22.9 115 188 21.3 72.6 10.3 2.00 8.20 1.00 5.06 0.95 2.38 0.27 1.58 0.22 429 417 11.46 36 52 0.7
23.3 202 375 37.2 133 17.0 3.76 12.1 1.14 4.82 0.79 2.03 0.25 1.67 0.22 791 780 10.92 71 87 0.8
PIR1
NBJ1
b
c
SIG
13.0 24.4 29.2 5.49 19.6 3.61 0.71 2.98 0.43 2.46 0.47 1.29 0.18 1.23 0.19 92 86 6.25 14 14 0.7
NBJ4
NPS
4.89 29.3 54.3 5.43 19.9 3.64 0.57 2.84 0.35 1.44 0.19 0.43 0.05 0.29 0.05 119 116 2.80 41 72 0.5 307 297 9
ATN2 d
10.3 63.1 121 11.9 39.6 6.74 0.87 5.32 0.61 2.62 0.36 0.78 0.08 0.48 0.07 254 249 5.00 50 94 0.4
PIR2
SBBS
19.0 57.8 110 12.1 48.2 8.19 2.30 6.87 0.87 3.89 0.65 1.82 0.23 1.50 0.21 255 245 9.17 27 28 0.9
ATB2 e
12.6 118 186 20.6 69.0 7.84 1.82 6.18 0.54 2.02 0.33 0.83 0.08 0.43 0.06 414 409 4.29 95 197 0.8
ATB3
ATB6
BEC1
BEC3
CAC0
CAC2
BPS
11.2 24.5 52.8 5.74 20.8 3.73 0.39 2.92 0.39 1.82 0.34 1.06 0.16 1.12 0.17 116 111 4.67 24 16 0.4
CAC1 f
18.4 57.0 105 11.2 41.9 6.70 1.81 5.45 0.68 3.28 0.59 1.86 0.25 1.86 0.27 238 229 8.79 26 22 0.9
24.1 111 268 21.8 76.9 11.4 1.40 9.08 1.13 5.85 0.99 2.81 0.35 2.30 0.32 513 500 13.75 36 35 0.4
30.5 72.5 161 19.0 70.1 11.5 2.46 8.93 1.13 5.97 1.11 2.94 0.37 2.36 0.33 360 345 14.21 24 22 0.7
50.4 77.7 188 22.4 86.0 15.4 2.48 11.6 1.67 9.43 1.81 4.74 0.62 3.78 0.51 426 404 22.56 18 15 0.6
43.6 70.6 154 17.8 75.5 13.8 3.12 11.7 1.62 8.94 1.65 4.40 0.57 3.64 0.47 368 347 21.29 16 14 0.8
7.20 15.6 24.0 2.49 8.80 1.59 1.29 1.50 0.23 1.33 0.26 0.82 0.12 0.82 0.12 59 55 3.70 15 14 2.6
PS
31.3 85.6 163 17.6 66.4 10.3 3.44 8.89 1.23 6.38 1.18 3.38 0.44 2.98 0.45 371 355 16.04 22 21 1.1
Piracaia Complex. Serra do Itaberaba Group. Nazaré Paulista Suite. Serra do Barro Branco Suite. Bragança Paulista Suite. Piracaia Suite, LREE = La to Gd and HREE = Tb to Lu (USGS - U.S. Geological Survey, 2013), Eu/Eu* = EuCN/(SmCN x GdCN)0.5 (McLennan, 1989). Median values of metamorphic (left) and igneous (right) rocks, the subscript CN implies chondrite normalized (Sun and McDonough, 1989).
414 mg kg−1, respectively), LREE enrichment (LREE/HREE = 27 and 95), moderate to large fractionation ratios [(La/Yb)CN = 28 and 197] and discrete Eu negative anomalies (Eu/Eu* = 0.9–0.8; Table 3). The significant differences in the ∑REE are noticeable in the chondritenormalized REE patterns (Fig. 5C). ATN2 contains amphibole, garnet, and zircon (≈8 wt.% of Fe-hornblende, ≈ 3 wt.% of andradite and ≈0.04 wt.% of Zrn, respectively) (Table 1b). The presence of amphibole in ATN2 could explain its less fractionated REECN pattern compared to that of PIR2 (≈4 wt.% of garnet and ≈0.05 wt.% of zircon). Despite the similarities in the LREE contents between amphiboles and feldspars, the chondrite-normalized REE patterns of Ca-amphiboles (like Fe-hornblende) are less fractionated (Bea, 1996). Overall, the results of Ragatky (1998) for samples of the same suite shows similar totals (∑REE = 284–328 mg kg−1), distribution patterns (LREE/HREE = 40–108, [La/Yb]CN = 44–205) and small Eu negative anomalies (Eu/Eu* = 0.7–0.8).
neighboring REE (Bea, 1996), and its positive Eu anomaly becomes compensated by a negative Eu/Eu* of other minerals (e.g., garnet), resulting in the absence of an anomaly for the whole-rock. The low ∑REE (59 mg kg−1) of granite CAC2 and its positive Eu anomaly (Eu/ Eu* = 2.6) were associated with its mineral composition, dominated by feldspar, quartz, plagioclase, some muscovite and amphibole, and the absence of accessory phases (Table 1b). CAC2 also presented relative depletion in Ti, Mg, Ca, and Zr and enrichment in Ba and Sr (Table 2 and Fig. A2 in Appendix A). 4.2.6. Piracaia Suite The total alkali vs. SiO2 (Fig. 4A) of monzonite CAC1, which represents the PS (Table 2), plots in the syenite field and exhibits a welldelineated shoshonitic affinity (Fig. 4B) as well as peraluminous nature (Fig. 4C). The REE signature of CAC1 is intermediary (∑REE = 371 mg kg−1, LREE/HREE = 22, [La/Yb]CN = 21) (Table 3) compared to other samples of the area, and there is an absence of an Eu anomaly (Eu/Eu* = 1.1).
4.2.5. Bragança Paulista Suite BPS samples comprise granitoids (ATB2, ATB3, ATB6, BEC1, CAC0, and CAC2) and one monzodiorite (BEC3) (Table 1b). At the refinement step of the mineral assemblages with MINSQ, the BEC3 significant MgO (4.34 wt.%), Fe2O3 (4.06 wt.%) and FeO (5.23 wt.%) contents were associated with the presence of biotite (12 wt.% eastonite) and amphiboles (14 wt.% pargasite and 15 wt.% of riebeckite). The ∑REE content of BPS samples ranged from 59 to 513 mg kg−1, with intermediary LREE enrichment (LREE/HREE = 15–36) and slightly fractionated patterns ([La/Yb]CN = 14–35) (Table 3). The REECN distribution patterns (Fig. 5D) exhibit negative Eu anomalies for samples ATB2 and ATB6 (Eu/Eu* = 0.4), BEC3 (Eu/Eu* = 0.6) and BEC1 (Eu/Eu* = 0.7). We associated the absence of an Eu anomaly (Eu/Eu* = 0.9) in monzogranite (ATB3) to its epidote content (Table 1b). When in granites, this phase hosts Eu preferably to its
5. Discussion The lithological framework of the six geological units of the Atibaia and Jaguari river subbasins comprises granitoids of monzogabbroic to granitic compositions, orthogneisses (dioritic to granitic compositions), paragneisses and local occurrences of quartzites. In the area, the most representative rocks are ortho and paragneisses of the PC and granitoids of the BPS. Moreover, at the basin-scale, the Atibaia predominantly comprises gneisses from the PC, whereas the Jaguari bedrocks include mostly granitoids from the Bragança Paulista Suite (Fig. 3). As rock groups, the igneous and metamorphic samples have similar median ΣREE (307 and 286 mg kg−1, respectively) (Table 3) and ΣLREE (297 and 271 mg kg−1, respectively). The median ΣHREE values 247
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 5. A to D - Chondrite-normalized REE patterns shown by geological units. A) Piracaia Complex and Serra do Itaberaba Group (PIR1). B) Nazaré Paulista Suite and Piracaia Suite (CAC1). C) Serra do Barro Branco Suite. D) Bragança Paulista Suite. Rock types and their acronyms are found in Tables 1a and 1b. *Sun and McDonough (1989).
Fig. 6. Chondrite-normalized REE interval obtained for in situ plagioclase grains analysis from orthogneiss ATB4. *Chondrite values of Sun and McDonough (1989).
Fig. 7. Chondrite-normalized REE interval obtained for in situ garnet grains analysis from orthogneiss ATB4. *Chondrite values of Sun and McDonough (1989).
are slightly higher in the metamorphic rocks compared to the igneous ones. All lithotypes of this study presented LREE enrichment relative to the HREE. To evaluate the role of the different rocks as REE carriers, the chondrite-normalized values of La vs. La/Yb, shown in Fig. 9, suggest discrete preferential LREE enrichment of the metamorphic (blue symbols) in comparison to the igneous (red symbols) rock groups. The monzogranites PIR2 (SBBS), NBJ4 and NBJ1 (NPS) and the gneisses CAC1.5 and BEC5 (PC) presented the highest REE fractionation (Table 3 and Fig. 9). Additionally, less evolved rocks such as the orthogneisses of dioritic (BEC4) and granodioritic (ATB4) compositions contain less ∑REE (and consequently LREE) than the gneisses derived from granites
(e.g., BEC5) (Table 3 and Fig. 9). The average ΣREE (308 mg kg−1) content of our samples is twice that value for the Upper Continental Crust (UCC ∼148 mg kg−1, Rudnick and Gao, 2003). The REE chondrite-normalized distributions (Fig. 10) demonstrates the LREE enrichment trend of our samples compared with the UCC, while the HREE split almost equally in the number of samples more enriched/depleted relatively to the UCC normalized values. Fig. 10 highlights the wide REE composition range of our rock samples and the area's lithological heterogeneity. Rock type, coexisting minerals and aluminum-saturation index (ASI), among other parameters, control the incorporation of REE in 248
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Fig. 10. REE chondrite-normalized distribution patterns of the Atibaia and Jaguari subbasins (AJSB) rock samples and the UCC values (Rudnick and Gao, 2003).
Fig. 8. Chondrite-normalized REE interval obtained for in situ plagioclase grains analysis from monzogranite NBJ1. *Chondrite values of Sun and McDonough (1989).
amphiboles contain less REE totals than its sodic variety, but both are slightly enriched in LREE over the HREE and marked by the absence of or a small Eu negative anomaly (Bea, 1996). According to our MINSQ estimates, Ca-amphiboles dominate over the sodic type in our samples. Two granitoids (ATN2 and BEC1) and one orthogneiss (BEC2) contain the higher proportion of Ca-amphiboles (≈8 wt.% of Fe-hornblende for the first two and ≈7 wt.% of Mghornblende in the later). These samples also have small negative Eu anomalies (Eu/Eu* = 0.7–0.9). Na-amphiboles occur in monzogranite (NBJ1) and monzodiorite (BEC3) (Table 1b). With MINSQ, we estimated riebeckite (≈1 wt.% in NBJ1 and ≈15 wt.% in BEC3) and pargasite (≈14 wt.% in BEC3). Naamphiboles present concave chondrite-normalized REE patterns from La to Dy followed by an increase from Er to Lu (Bea, 1996). The REE normalized pattern of sample BEC3 somehow fits into such a description and exhibits the highest ΣHREE (≈23 mg kg−1) among all the samples, probably because of its large proportion of Na-amphiboles. Accessory minerals contain a significant proportion of the overall REE budget of felsic rocks, either as REE phases (monazite, xenotime, and allanite) or as minor and trace elements (apatite, titanite, garnet, and zircon), only to cite some typical examples. Allanite occurs in gneisses BEC5 (≈0.04 wt.%) and CAC1.5 (≈0.1 wt.%) (Table 1a). Among all samples, CAC1.5 has the highest ∑REE (791 mg kg−1), and allanite probably hosts part of this budget, besides titanite (≈4 wt.%). Allanite was only identified in metamorphic rocks of the PC, which composes the area's dominant geological unit (Table A1 - Appendix A), as seen principally for the Atibaia subbasin (Fig. 3). Apatite may contain a variable proportion of REE (Roeder et al., 1987), while the fractionation of the REE and the Eu anomaly in this mineral depend on the rock's aluminosity. According to Bea (1996), apatites from peraluminous and metaluminous lithotypes differ in the REE totals and Eu anomalies. The first group contains more REE than metaluminous apatite and presents pronounced negative Eu anomaly (e.g., Eu/Eu* ≈ 0.1), while the last carry proportionally more HREE and presents less intense Eu anomaly (e.g., Eu/Eu* ≈ 0.7). Apatite was identified in samples of three geological units (PC, NPS, and granitoids of BPS). The granitoids NBJ1, ATB6, BEC1, BEC3, and CAC0 have 0.1 to 1 wt.% of apatite, based on MINSQ estimates (Table 1b). The first two, NBJ1 and ATB6, are peraluminous granites and display significantly fractionated REE patterns, as already discussed, and negative Eu anomalies (Eu/Eu* = 0.5 and 0.4, respectively). On their turn, the metaluminous granitoids BEC1, BEC3 and CAC0 contain ≈1 wt.% of apatite, are less fractionated ([La/Yb]CN = 14–22) and presented
Fig. 9. Relative REE enrichments expressed as LaCN vs. (La/Yb)CN for metamorphic (blue symbols) and igneous (red symbols) rock samples of the Atibaia and Jaguari subbasins. Symbols: Square = Piracaia Complex, Circle = Serra do Itaberaba Group, Triangle = Nazaré Paulista Suite, Diamond = Serra do Barro Branco Suite, Star = Bragança Paulista Suite and Pentagon = Piracaia Suite. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article).
feldspars, the LREE contents of which are typically moderate to low, while those of HREE are low to very low (Bea, 1996). Moreover, feldspars of highly fractionated peraluminous granites contain relatively low REE totals (Bea et al., 1994), as found in samples NBJ1 ([La/ Yb]CN = 72; ∑REE = 119 mg kg−1) and NBJ4 ([La/Yb]CN = 94; ∑REE = 254 mg kg−1) (Table 3 and Fig. 4C). The LREE enrichment, the strong LREE-HREE fractionation and the positive Eu anomalies were confirmed by in situ analysis of individual plagioclase grains of orthogneiss ATB4 and monzogranite NBJ1 (Figs. 6 and 8, respectively). The comparison of Fig. 8 with the plot of the whole-rock sample (Fig. 5B) shows that the plagioclase REE content in sample NBJ1 contributes significantly to the relative LREE enrichment seen for the NPS granitic rocks. The REE signatures of these particular plagioclase grains are also expected to occur in the remaining samples of similar genesis (Bea, 1996) and accordingly produce the observed whole-rock REE patterns. Ultimately, minerals more enriched in REE than feldspars control the total contents of these elements in a specific rock. Among the major minerals, amphibole is more enriched in REE than feldspar, and Ca249
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
Among the HREE enriched minerals (Bea, 1996) the more recurrently present in our samples were Na-amphibole, garnet, and zircon. The first two are known for their low resistance during weathering reactions (Velbel, 1985; Price et al., 2005b; Velbel and Price, 2007), while zircon is recognized for its stability during the rock cycle (Balan et al., 2001; Hanchar and van Westrenen, 2007; Sanematsu et al., 2015, 2016). Published REE data obtained of fourteen water samples from Atibaia and Jaguari rivers (Campos and Enzweiler, 2016), show (La/ Yb) shale-normalized (McLennan, 1989) values between 0.94 and 1.82, with 64% of samples having LREE enriched patterns. The shale-normalized La/Yb ratios of the rock samples of this study fall within 1.2 and 20.3, with a median value of 2.3, showing a reasonable agreement between rock and water samples, despite the low solubility of the REE under the circumneutral pH and low salinity local prevailing conditions. During the water-rock interaction, newly formed phases or secondary minerals tend to retain most of the released REE budget. The REE behavior as nonmobile elements (Gaillardet et al., 2014) and their fractionation between solid and liquid phases, according to media composition and specific complexation affinity to ligands, can result in complex scenarios. Besides, punctual and diffuse anthropogenic inputs may also affect the dissolved REE measured values.
smaller Eu negative anomalies (Eu/Eu* = 0.7, 0.6 and 0.8, respectively). Garnet was frequently identified both in metamorphic and igneous rock samples. This mineral concentrates HREE (Hönig et al., 2014 and references therein), reaching up to ≈ 4 wt.% in igneous (Wang et al., 2003) and ≈13 wt.% in metamorphic garnets (Grew et al., 2010). Our MINSQ estimates of garnet contents in the metamorphic and igneous rocks showed a close association with the HREE totals in the respective samples. Despite the relative lower abundances of garnets in the metamorphic rock samples (ATB4, ATB5, BEC2, BEC4 and BEC5 with approximately 4, 0.4, 0.2, 0.4 and 0.05 wt.%, respectively) compared to the igneous ones (NBJ1, NBJ4, ATN2, PIR2 and ATB3 with approximately 0.02, 1, 3, 4 and 3 wt.%, respectively), the first exhibited higher HREE contents (∑HREE ≈ 12, 18, 18, 11 and 11 mg kg−1, respectively; Table 3). The chondrite-normalized patterns of orthogneiss ATB4 garnet grains (Fig. 7) confirm the HREE content, and the contribution of this phase to the whole-rock negative Eu anomaly observed for this sample (Eu/Eu* = 0.8) (Table 3). All metamorphic garnet-bearing rocks of this study presented Eu anomalies (Eu/Eu* = 0.4 to 0.8; see Table 3), suggesting that despite the relatively low abundance of garnet in these samples, this mineral influences the respective rocks’ REE patterns. Zircon is a recurrent accessory phase identified in our rock samples. Zircon crystals can incorporate a significant amount of REE depending on growth conditions (pressure, temperature and melt/fluid composition) (Hanchar and van Westrenen, 2007). The HREE smaller ionic sizes have higher affinity with the crystallographic sites in zircon, and, as a result, the REECN patterns of this mineral are progressively enriched from Gd to Yb and display a moderate negative Eu anomaly (Bea, 1996). We also suggest zircon as a potential host phase for the HREE in selected samples. Despite the low zircon content (0.03 wt.% estimated using MINSQ) in quartzite PIR1, we suggest this phase bears most of the ∑HREE (≈6 mg kg−1; Table 3) measured for this sample and contributes to its whole-rock Eu anomaly (Eu/Eu* = 0.7). Similarly, from the amounts of Zrn (≈0.07, 0.07 and 0.06 wt.%, respectively; Table 1a), negative Eu anomalies (Eu/Eu* = 0.7, 0.7 and 0.8, respectively; Table 3) and little or absent garnet in metamorphic rocks BEC2, BEC5, and CAC1.5, we infer that zircon has a crucial role in the REE signatures of these samples. To extend the results of this study and preview how the mineral and chemical composition of the geological framework will influence the hydrogeochemistry of the REE, we should consider many different agents. In addition to the REE distributions in the lithotypes, their distinct hosting minerals and weathering susceptibility under the area's chemical, physical and biological processes comprehend the primary mechanisms behind the REE mobility. Climatic conditions, anthropogenic inputs, and tectonic dynamics compose additional forces that affect the processes of the “weathering engine” (Anderson et al., 2004). Here, we focus on the general REE trends of the rock samples, their most significant budget of hosting minerals and their weathering susceptibility. The rocks of the area have REE signatures typical of the evolved lithosphere, both in REE total content and LREE enrichment. Above, we established some plausible relationships between these features and their respective principal and accessory minerals. A significant fraction of the whole-rock REE content associates to some of the most abundant minerals (e.g., feldspar and plagioclase), and some samples contain Ca-amphibole and minor contents of apatite, allanite and limited titanite. These minerals are all enriched in LREE compared to HREE (Bea, 1996) and are among the more susceptible to weathering (Schott et al., 1981; Banfield and Eggleton, 1989; Velbel, 1999; Price et al., 2005a; White and Buss, 2014). As a result, we could expect weathering products enriched in the LREE. However, at least some fractionation of the REE will occur as a consequence of the different affinities along the series regarding incorporation by secondary minerals and complexation by solution ligands.
6. Conclusions Gneisses of the Piracaia Complex and granitoids of the Bragança Paulista Suite are the dominant lithotypes among the six principal geological units of the Atibaia and Jaguari subbasins, respectively. The REE distribution patterns of all the samples presented enrichment in the LREE compared to the HREE. Locally, the metamorphic rocks of the Atibaia subbasin contain higher REE totals and show distribution patterns relatively more enriched and fractionated between LREE than HREE in comparison with the igneous rocks of the Jaguari subbasin. Negative Eu anomalies, observed for many samples, probably reflect some more prominent Eu/Eu* signatures of individual minerals. The main REE hosting phases, estimated on the basis of their abundance in the rock samples and as potential REE carriers, are Caamphibole, plagioclase, K-feldspar and minor amounts of apatite, allanite, and titanite for the LREE and Na-amphibole, garnet and zircon for the HREE. Considering only the REE budget of the two subbasins and the weathering susceptibility of the minerals retaining these elements, the metamorphic rocks of the Piracaia Complex should contribute with a more LREE-enriched signature to the Atibaia river water in comparison to that of the Jaguari river, which watershed is influenced by the Bragança Paulista granitoids. Acknowledgments BCM acknowledges a doctorate scholarship from the Coordination for the Improvement of Higher Education Personnel (CAPES). JE acknowledges the support from the National Council for Scientific and Technological Development (CNPq, grant number 312507/2013-5) and São Paulo Research Foundation (FAPESP, grant number 2012/050242). The authors are grateful to J.S. Armstrong-Altrin and two anonymous reviewers for their constructive comments. Appendix A. Supplementary data Supplementary data to this article can be found online at https:// doi.org/10.1016/j.jsames.2019.01.017. References Amonette, J.E., Templeton, J.C., 1998. Improvements to the quantitative assay of nonrefractory minerals for Fe(II) and total Fe using 1,10-Phenanthroline. Clay Clay
250
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
escala 1:50000 (in Portuguese) In: Proceedings of the XXXVIII Brazilian Geological Congress. Balneário Camboriú, Brazil, pp. 79–81. Hagedorn, B., Cartwright, I., Raveggi, M., Maas, R., 2011. Rare earth element and strontium geochemistry of the Australian Victorian Alps drainage system: Evaluating the dominance of carbonate vs. aluminosilicate weathering under varying runoff. Chem. Geol. 284, 105–126. https://doi.org/10.1016/j.chemgeo.2011.02.013. Hanchar, J.M., van Westrenen, W., 2007. Rare Earth element behavior in zircon-melt systems. Elements 3, 37–42. https://doi.org/10.2113/gselements.3.1.37. Hartmann, J., 2009. Bicarbonate-fluxes and CO2-consumption by chemical weathering on the Japanese Archipelago - Application of a multi-lithological model framework. Chem. Geol. 265, 237–271. https://doi.org/10.1016/j.chemgeo.2009.03.024. Heilbron, M., Machado, N., 2003. Timing of terrane accretion in the NeoproterozoicEopaleozoic Ribeira Orogen (SE Brazil). Precambrian Res. 125, 87–112. https://doi. org/10.1016/S0301-9268(03)00082-2. Heilbron, M., Soares, A.C.P., Campos Neto, M.C., Silva, L.C., Trouw, R.A.J., Janasi, V.A., 2004. Mantiqueira province. In: Neto, V.M., Bartorelli, A., Carneiro, C.D.R., Neves, B.B.B. (Eds.), Geologia do Continente Sul-Americano: Evolução da Obra de Fernando Flávio Marques de Almeida (in Portuguese). Beca, São Paulo, 8587256459, pp. 203–234. Henderson, P., 1984. General geochemical properties and abundances of the rare earth elements. In: Henderson, P. (Ed.), Developments in Geochemistry 2: Rare Earth Element Geochemistry. Elsevier Science Publishers, New York, 9781483289779, pp. 1–32. Herrmann, W., Berry, R.F., 2002. MINSQ - a least squares spreadsheet method for calculating mineral proportions from whole rock major element analyses. Geochem. Explor. Environ. Anal. 2 (4), 361–368. https://doi.org/10.1144/1467-787302-010. Hönig, S., Čopjaková, R., Škoda, R., Novák, M., Dolejš, D., Leichmann, J., Galiová, M.V., 2014. Garnet as a major carrier of the Y and REE in the granitic rocks: an example from the layered anorogenic granite in the Brno Batholith, Czech Republic. Am. Mineral. 99, 1922–1941. https://dx.doi.org/10.2138/am-2014-4728. Hoyle, J., Elderfield, H., Gledhill, A., Greaves, M., 1984. The behaviour of the rare earth elements during mixing of river and seawaters. Geochem. Cosmochim. Acta 48 (1), 143–149. https://doi.org/10.1016/0016-7037(84)90356-9. Iwata, S.A., 1995. Pegmatitos Graníticos da região de Socorro-SP (in Portuguese). Master's dissertation. Geosciences Institute, University of São Paulo, São Paulo, Brazil 122 pp. Janasi, V.A., 1999. Petrogênese de granitos crustais na Nappe de empurrão SocorroGuaxupé (SP-MG): Uma contribuição da geoquímica elemental e isotópica (in Portuguese). Habilitation thesis. Geosciences Institute, University of São Paulo, São Paulo, Brazil 304 pp. Jochum, K.P., Weis, U., Stoll, B., Kuzmin, D., Yang, Q., Raczek, I., Jacob, D.E., Stracke, A., Birbaum, K., Frick, D.A., Günther, D., Enzweiler, J., 2011. Determination of reference values for NIST SRM 610-617 glasses following ISO guidelines. Geostand. Geoanal. Res. 35 (4), 397–429. https://doi.org/10.1111/j.1751-908X.2011.00120.x. Juliani, C., Beljavskis, P., 1995. Revisão da litoestratigrafia da Faixa São Roque/Serra do Itaberaba - SP (in Portuguese). Rev. Institut. Geol. 16, 33–58. https://dx.doi.org/10. 5935/0100-929X.19950003. Keasler, K.M., Loveland, W.D., 1982. Rare earth elemental concentrations in some Pacific Northwest rivers. Earth Planet. Sci. Lett. 61, 68–72. https://doi.org/10.1016/0012821X(82)90039-5. Machado, N., Valladares, C., Heilbron, M., Valeriano, C., 1996. U-Pb geochronology of the central Ribeira belt (Brazil) and implications for the evolution of the Brazilian Orogeny. Precambrian Res. 79 (3–4), 347–361. https://doi.org/10.1016/03019268(95)00103-4. McLennan, S.M., 1989. Rare earth elements in sedimentary rocks: influence of provenance and sedimentary processes. Rev. Mineral. 21 (1), 169–200. Middlemost, E.A.K., 1994. Naming materials in the magma/igneous rock system. Earth Sci. Rev. 37 (3–4), 215–224. https://doi.org/10.1016/0012-8252(94)90029-9. Millot, R., Gaillardet, J., Dupré, B., Allègre, C.J., 2002. The global control of silicate weathering rates and the coupling with physical erosion: new insights from rivers of the Canadian shield. Earth Planet. Sci. Lett. 196, 83–98. https://doi.org/10.1016/S0012-821X(01)00599-4. Morais, S.M., 1999a. Programa Levantamentos Geológicos Básicos do Brasil: Projeto de Integração Geológica da Folha Campinas (SF.23-Y-A), escala 1:250000 (in Portuguese). Brazilian Geological Survey (CPRM), São Paulo, Brazil 40 pp. Morais, S.M., 1999b. Programa Levantamentos Geológicos Básicos do Brasil: Projeto de Integração Geológica da Folha Guaratinguetá (SF.23-Y-B), escala 1:250000 (in Portuguese). Brazilian Geological Survey (CPRM), São Paulo, Brazil 41 pp. Morais, S.M., 1999c. Programa Levantamentos Geológicos Básicos do Brasil: Projeto de Integração Geológica da Folha Santos (SF.23-Y-D), escala 1:250000 (in Portuguese). Brazilian Geological Survey (CPRM), São Paulo, Brazil 47 pp. Navarro, M.S., Tonetto, E.M., Oliveira, E.P., 2015. LA-SF-ICP-MS U-Pb zircon dating at University of Campinas, Brazil. In: Proceedings of the IX International Conference on the Analysis of Geological and Environmental Materials (Geoanalysis 2015). Leoben, Austria, pp. 86. Négrel, P., 2006. Water-granite interaction: clues from strontium, neodymium and rare earth elements in soil and waters. Appl. Geochem. 21, 1432–1454. https://doi.org/ 10.1016-/j.apgeochem.2006.04.007. Nesbitt, H.W., 1979. Mobility and fractionation of rare earth elements during weathering of a granodiorite. Nature 279, 206–210. https://doi.org/10.1038/279206a0. Pattan, J.N., Pearce, N.J.G., Mislankar, P.G., 2005. Constraints in using cerium-anomaly of bulk sediments as an indicator of paleo bottom water redox environment: a case study from the Central Indian Ocean Basin. Chem. Geol. 221, 260–278. https://doi. org/-10.1016/j.chemgeo.2005.06.009. Peccerillo, A., Taylor, S.R., 1976. Geochemistry of eocene calc-alkaline volcanic rocks from the Kastamonu area, Northern Turkey. Contrib. Mineral. Petrol. 58 (1), 63–81. https://doi.org/10.1007/BF00384745.
Miner. 46 (1), 51–62. https://doi.org/10.1346/CCMN.1998.0460106. Anderson, S.P., Blum, J., Brantley, S.L., Chadwick, O., Chorover, J., Derry, L.A., Drever, J.I., Hering, J.G., Kirchner, J.W., Kump, L.R., 2004. Proposed initiative would study Earth's weathering engine. Eos, Transactions, American Geophysical Union 85 (28), 265–272. https://doi.org/10.1029/2004EO280001. Armstrong-Altrin, J.S., Machain-Castillo, M.L., Rosales-Hoz, L., Carranza-Edwards, A., Sanchez-Cabeza, J.-A., Ruíz-Fernández, A.C., 2015. Provenance and depositional history of continental slope sediments in the Southwestern Gulf of Mexico unraveled by geochemical analysis. Cont. Shelf Res. 95, 15–26. https://doi.org/10.1016/j.csr. 2015.01.003. Armstrong-Altrin, J.S., Lee, Y.I., Kasper-Zubillaga, J.J., Trejo-Ramírez, E., 2017. Mineralogy and geochemistry of sands along the Manzanillo and El Carrizal beach areas, southern Mexico: implications for palaeoweathering, provenance and tectonic setting. Geol. J. 52, 559–582. https://doi.org/10.1002/gj.2792. Armstrong-Altrin, J.S., Ramos-Vázquez, M.A., Zavala-León, A.C., Montiel-García, P.C., 2018. Provenance discrimination between Atasta and Alvarado beach sands, western Gulf of Mexico, Mexico: Constraints from detrital zircon chemistry and U-Pb geochronology. Geol. J. 1–25. https://doi.org/10.1002/gj.3122. Artur, A.C., Wernick, E., Hôrmann, P.K., Weber-Diefenbach, K., 1993. Associações Plutônicas do Complexo Granitóide Socorro (Estados de São Paulo e Minas Gerais, SE Brasil) (in Portuguese). Braz. J. Geol. 23 (3), 265–273. Aubert, D., Stille, P., Probst, A., 2001. REE fractionation during granite weathering and removal by waters and suspended loads: Sr and Nd isotopic evidence. Geochem. Cosmochim. Acta 65 (3), 387–406. https://doi.org/10.1016/S0016-7037(00) 00546-9. Balan, E., Trocellier, P., Jupille, J., Fritsch, E., Muller, J.-P., Calas, G., 2001. Surface chemistry of weathered zircons. Chem. Geol. 181, 13–22. https://doi.org/10.1016/ S0009-2541(01)00271-6. Banfield, J.F., Eggleton, R.A., 1989. Apatite replacement and rare earth mobilization, fractionation, and fixation during weathering. Clay Clay Miner. 37 (2), 113–127. https://doi.org/10.1346/CCMN.1989.0370202. Basei, M.A.S., Frimmel, H.E., Nutman, A.P., Preciozzi, F., 2008. West Gondwana amalgamation based on detrital zircon ages from Neoproterozoic Ribeira and Dom Feliciano belts of South America and comparison with coeval sequences from SW Africa. Geological Society of London Special Publications 294, 239–256. https://doi. org/10.1144/SP294.1-3. Bea, F., Pereira, M.D., Corretgé, L.G., Fershtater, G.B., 1994. Differentiation of strongly peraluminous, perphosphorus granites. The Pedrobernardo pluton, central Spain. Geochem. Cosmochim. Acta 58 (12), 2609–2627. https://doi.org/10.1016/00167037(94)90132-5. Bea, F., 1996. Residence of REE, Y, Th and U in granites and crustal protoliths; implications for the chemistry of crustal melts. J. Petrol. 37 (3), 521–552. https://doi. org/10.-1093/petrology/37.3.521. Braga, I.F., 2002. Análise Deformacional de Rochas Infracrustais da Região de Cristina e Itajubá-MG (in Portuguese). PhD thesis. Geosciences Institute, UNESP Universidade Estadual Paulista “Júlio de Mesquita Filho”, São Paulo, Brazil 216 pp. Campos, F.F., Enzweiler, J., 2016. Anthropogenic gadolinium anomalies and rare earth elements in the water of Atibaia River and Anhumas Creek, Southeast Brazil. Environ. Monit. Assess. 188, 281. https://doi.org/10.1007/s10661-016-5282-7. Campos Neto, M.C., Figueiredo, M.C.H., Basei, M.A.S., Alves, F.R., 1984. Os granitóides da região de Bragança Paulista (SP) (in Portuguese) In: Proceedings of the XXXIII Brazilian Geological Congress, Rio de Janeiro, Brazil, pp. 1809–1822. CBH-PCJ - Committee of Hydrographic Basins of Piracicaba, 2016. Capivari and Jundiaí rivers. Available at: www.agenciapcj.org.br/docs/relatorios/relatorio-situacao2015.pdf, Accessed date: 12 February 2017 (in Portuguese). Cotta, A.J.B., Enzweiler, J., 2008. Certificate of analysis of the reference material BRP-1 (Basalt Ribeirão Preto). Geostand. Geoanal. Res. 32 (2), 231–235. https://doi.org/10. 1111/j.1751-908X.2008.00894.x. Cotta, A.J.B., Enzweiler, J., 2012. Classical and new procedures of whole rock dissolution for trace element determination by ICP-MS. Geostand. Geoanal. Res. 36 (1), 27–50. https://doi.org/10.1111/j.1751-908X.2011.00115.x. Elderfield, H., Upstill-Goddard, R., Sholkovitz, E.R., 1990. The rare earth elements in rivers, estuaries, and coastal seas and their significance to the composition of ocean waters. Geochem. Cosmochim. Acta 54, 971–991. https://doi.org/10.1016/00167037(90)90432-K. Enzweiler, J., Vendemiatto, M.A., 2013. Estimativa da incerteza de medição na determinação de elementos maiores e menores em rochas silicáticas por espectrometria de fluorescência de raios X com resultados de controle de qualidade (in Portuguese). Geochim. Bras. 27 (2), 152–160. https://doi.org/10.5327/Z010298002013000200-07. Gaillardet, J., Viers, J., Dupré, B., 2014. Trace elements in river waters. In: Turekian, K., Holland, H. (Eds.), Treatise on Geochemistry. Elsevier Science Publishers, New York, 9780080983004, pp. 195–232. Gengo, R.M., 2014. Petrologia de Ortognaisses e Granitóides do Domínio Socorro, Nappe Socorro-Guaxupé, seção Extrema-Camanducaia (in Portuguese). Master’s dissertation. Geosciences Institute, University of São Paulo, São Paulo, Brazil 110 pp. Goldschmidt, V.M., 1937. The principles of distribution of chemical elements in minerals and rocks. J. Chem. Soc. 655–673. https://doi.org/10.1039/JR93700006-55. Goldstein, S.J., Jacobsen, S.B., 1988. Rare earth elements in river waters. Earth Planet. Sci. Lett. 89, 35–47. https://doi.org/10.1016/0012-821X(88)90031-3. Grew, E.S., Marsh, J.H., Yates, M.G., Lazic, B., Armbruster, T., Locock, A., Bell, S.W., Dyar, M.D., Bernhardt, H.-J., Medenbach, O., 2010. Menzerite-(Y), a new species, {(Y,REE)(Ca,Fe2+)2}[(Mg,Fe2+)(Fe3+,Al)](Si3)O12, from a felsic granulite, Parry Sound, Ontario, and a new garnet end-member, {Y2Ca}[Mg2](Si3)O12. Can. Mineral. 48 (5), 1171–1193. https://dx.doi.org/10.3749/canmin.48.5.1171. Hackspacher, P.C., Godoy, A.M., Oliveira, M.A.F., 1994. Geologia da Folha Cabreúva - SP,
251
Journal of South American Earth Sciences 91 (2019) 239–252
B.C. Mortatti, et al.
in the Ribeira Belt (southeastern Brazil): The Pirapora do Bom Jesus ophiolitic complex. Episodes 24 (4), 245–251. Tassinari, C.C.G., Babinski, M., Nutman, A.P., 2004. A idade e a natureza da Fonte do Granito do Moinho, Faixa Ribeira, Sudeste do Estado de São Paulo (in Portuguese). Geol. Usp. Série Científica 4 (1), 91–100. https://dx.doi.org/10.5327/S1519874x2004000100006. Taylor, S.R., McLennan, S.M., 1985. The Continental Crust: its Composition and Evolution. Blackwell Scientific Publications, Oxford, 978-0632011483pp. 312. Taylor, S.R., McLennan, S.M., 1995. The geochemical evolution of the continental crust. Rev. Geophys. 33 (2), 241–265. https://doi.org/10.1029/95RG00262. Tricca, A., Stille, P., Steinmann, M., Kiefel, B., Samuel, J., Eikenberg, J., 1999. Rare earth elements and Sr and Nd isotopic compositions of dissolved and suspended loads from small river systems in the Vosges mountains (France), the river Rhine and groundwater. Chem. Geol. 160, 139–158. https://doi.org/10.1016/S0009-2541(99) 00065-0. Ulbrich, H.H., Vlach, S.R.F., Demaiffe, D., Ulbrich, M.N.C., 2005. Structure and origin of the Poços de Caldas alkaline massif, SE Brazil. In: In: Comin-Chiaramonti, P., Gomes, C.B. (Eds.), Mesozoic to Cenozoic Alkaline Magmatism in the Brazilian Platform, vol. 367. Publisher: FAPESP, São Paulo, 85-314-0903-9, pp. 418. USGS - U.S. Geological Survey, 2013. 2011 Minerals Yearbook - Rare Earths (Advance Release). U.S. Geological Survey, Reston, pp. 13. Available at: https://minerals.usgs. gov/min-erals/pubs/commodity/rare_earths/myb1-2011-raree.pdf, Accessed date: 4 March 2017. Velbel, M.A., 1985. Geochemical mass balances and weathering rates in forested watersheds of the southern Blue Ridge. Am. J. Sci. 285, 904–930. https://doi.org/10.2475/ ajs.285.10.904. Velbel, M.A., 1999. Bond strength and the relative weathering rates of simple orthosilicates. Am. J. Sci. 299 (7), 679–696. https://doi.org/10.2475/ajs.299.7-9.679. Velbel, M.A., Price, J.R., 2007. Solute geochemical mass-balances and mineral weathering rates in small watersheds: Methodology, recent advances, and future directions. Appl. Geochem. 22, 1682–1700. https://doi.org/10.1016/j.apgeochem.2-007.03.029. Vendemiatto, M.A., Enzweiler, J., 2001. Routine control of accuracy in silicate rock analysis by X-ray fluorescence spectrometry. Geostand. Newsl. J. Geostand. Geoanal. 25 (2–3), 283–291. https://doi.org/10.1111/j.1751-908X.2001.tb00604.x. Vlach, S.R.F., 1985. Geologia, Petrografia e Geocronologia das regiões meridional e oriental do Complexo de Morungaba, SP. Master's dissertation. Geosciences Institute, University of São Paulo, São Paulo, Brazil, pp. 253. Wang, R.C., Fontan, F., Chen, X.M., Hu, H., Liu, C.S., Xu, S.J., Parseval, P., 2003. Accessory minerals in the Xihuashan Y-enriched granitic complex, southern China: A record of magmatic and hydrothermal stages of evolution. Can. Mineral. 41, 727–748. https://dx.doi.org/10.2113/gscanmin.41.3.727. Warne, A.G., Meade, R.H., White, W.A., Guevara, E.H., Gibeaut, J., Smyth, R.C., Aslan, A., Tremblay, T., 2002. Regional controls on geomorphology, hydrology, and ecosystem integrity in the Orinoco Delta, Venezuela. Geomorphology 44, 273–307. https://doi. org/10.1016/S0169-555X(01)00179-9. Webmineral database - Available at: https://www.webmineral.com. Accessed in: 11/06/ 2017. Wernick, E., Artur, A.C., Hörmann, P.K., Weber-Diefenbach, K., Fahl, F.C., 1997. O magmatismo alcalino potássico Piracaia, SP (SE Brasil): Aspectos composicionais e evolutivos (in Portuguese). Braz. J. Geol. 27 (1), 53–66. White, A.F., Buss, H.L., 2014. Natural weathering rates of silicate minerals. In: Turekian, K., Holland, H. (Eds.), Treatise on Geochemistry. Elsevier Science Publishers, New York, 9780080983004, pp. 195–232. Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals. Am. Mineral. 95, 185–187. https://doi.org/10.2138/am.2010.3371. Xu, Z., Han, G., 2009. Rare earth elements (REE) of dissolved and suspended loads in the Xijiang River, South China. Appl. Geochem. 24, 1803–1816. https://doi.org/10. 1016/j.-apgeochem.2009.06.001.
Peloggia, A.U.G., 1990. A Faixa Alto Rio Grande na região de Amparo (SP) (in Portuguese). Master's dissertation. Geosciences Institute, University of São Paulo, São Paulo, Brazil 124 pp. Price, J.R., Velbel, M.A., Patino, L.C., 2005a. Allanite and epidote weathering at the Coweeta Hydrologic Laboratory, western North Carolina, U.S.A. Am. Mineral. 90, 101–114. https://doi.org/10.2138/am.2005.1444. Price, J.R., Velbel, M.A., Patino, L.C., 2005b. Rates and timescales of clay-mineral formation by weathering in saprolitic regoliths of the southern Appalachians from geochemical mass balance. Geol. Soc. Am. Bull. 117, 783–794. https://doi.org/10. 1130/B25547.1. Ragatky, C.D., 1998. Contribuição à geoquímica e geocronologia do Domínio São Roque e da Nappe de empurrão Socorro-Guaxupé na região de Igaratá e Piracaia, SP (in Portuguese). PhD thesis. Geosciences Institute, University of São Paulo, São Paulo, Brazil 130 pp. Roeder, P.L., MacArthur, D., Ma, X.-P., Palmer, G.R., 1987. Cathodoluminescence and microprobe study of rare-earth elements in apatite. Am. Mineral. 72, 801–811. Rudnick, R.L., Gao, S., 2003. Composition of the continental crust. In: Turekian, K., Holland, H. (Eds.), Treatise on Geochemistry. Elsevier-Pergamon, Oxford, pp. 1–64. https://doi.org/10.1016/B0-08-043751-6/030-4. SABESP - Basic Sanitation Company of the State of São Paulo, 2017 http://www.sabesp. com.br, Accessed 15 March 2017, (in Portuguese). Sachs, L.L.B., 1999. Programa Levantamentos Geológicos Básicos do Brasil: Projeto de Integração Geológica da Folha São Paulo (SF.23-Y-C), escala 1:250000 (in Portuguese). Brazilian Geological Survey (CPRM), São Paulo, Brazil 36 pp. Sanematsu, K., Kon, Y., Imai, A., 2015. Influence of phosphate on mobility and adsorption of REEs during weathering of granites in Thailand. J. Asian Earth Sci. 111, 14–30. https://dx.doi.org/10.1016/j.jseaes.2015.05.018. Sanematsu, K., Ejima, T., Kon, Y., Manaka, T., Zaw, K., Morita, S., Seo, Y., 2016. Fractionation of rare-earth elements during magmatic differentiation and weathering of calc-alkaline granites in southern Myanmar. Mineral. Mag. 80 (1), 77–102. https:// doi.org/10.1180/minmag.2016.080.053. Schott, J., Berner, R.A., Sjöberg, E.L., 1981. Mechanism of pyroxene and amphibole weathering. I. Experimental studies of iron-free minerals. Geochem. Cosmochim. Acta 45, 2133–2135. https://doi.org/10.1016/0016-7037(81)90065-X. Shand, J.S., 1943. Eruptive Rocks. Thomas Murby & Co., London 444 pp. Smith, C., Liu, X.-M., 2018. Spatial and temporal distribution of rare earth elements in the Neuse River, North Carolina. Chem. Geol. 488, 34–43. https://doi.org/10.101-6/j. chemgeo.2018.04.003. Steinmann, M., Stille, P., 2008. Controls on transport and fractionation of the rare earth elements in stream water of a mixed basaltic-granitic catchment basin (Massif Central, France). Chem. Geol. 254, 1–18. https://doi.org/10.1016/j.chemgeo.2008. 04.004. Streckeisen, A.L., 1974. Classification and nomenclature of Plutonic rocks: recommendations of the IUGS subcommission on the systematics of Igneous rocks. Geol. Rundsch. 63 (2), 773–786. https://doi.org/10.1007/BF01820841. Sun, S-s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts: implications for mantle composition and processes 42. Geological Society of London Special Publications, pp. 313–345. https://doi.org/10.1144/GSL.SP.1989.042.01.19. Tapia-Fernandez, H.J., Armstrong-Altrin, J.S., Selvaraj, K., 2017. Geochemistry and U-Pb geochronology of detrital zircons in the Brujas beach sands, Campeche, Southwestern Gulf of Mexico, Mexico. J. S. Am. Earth Sci. 76, 346–361. https://doi.org/10.1016/j. jsames.2017.04.003. Tardy, Y., Bustillo, V., Roquin, C., Mortatti, J., Victoria, R., 2005. The Amazon. Biogeochemistry applied to river basin management. Part I - Hydro-climatology, hydrograph separation, mass transfer balances, stable isotopes, and modelling. Appl. Geochem. 20, 1746–1829. https://doi.org/10.1016/j.apgeochem.2005.06.001. Tassinari, C.C.G., Munhá, J.M.U., Ribeiro, A., Correia, C.T., 2001. Neoproterozoic oceans
252